Equivalent Delta-V Per Orbit of Gravitational Perturbations∗

Total Page:16

File Type:pdf, Size:1020Kb

Equivalent Delta-V Per Orbit of Gravitational Perturbations∗ Equivalent delta-v per orbit of gravitational perturbations∗ Martin Laray Polytechnic University of Madrid – UPM, 28040 Madrid, Spain Accepted: January 5, 2016 Nomenclature a = semi-major axis (orbital element); length unit C= subindex for a term derived from the Coriolis effect c= subindex for a term derived from the centrifugal forces D= subindex for disturbing function e = eccentricity (orbital element); dimensionless f = true anomaly; rad G = angular momentum vector per unit of mass; (length unit)2=time unit I = inclination (orbital element); rad J = impulse per unit of mass (delta-v); length unit/time unit J2 = zonal harmonic coefficient of the second degree; dimensionless K= subindex for Keplerian k = unit vector in the z direction; dimensionless M = mean anomaly; rad N = reference system’s rotation rate; rad/time unit N = modulus of N; rad/time unit n = orbit mean motion n = pµ/a3; rad/time unit p = conic parameter p = aη2; length unit ∗Journal of Guidance, Control, and Dynamics, in press, http://arc.aiaa.org/loi/jgcd Copyright c 2016 by the American Institute of Aeronautics and Astronautics, Inc. All rights re- served. yResearcher. Space Dynamics Group, ETSI Aeronauticos,´ Pz. Cardenal Cisneros 3, Associate Fellow AIAA. 1 q = auxiliary symbol of order 1; dimensionless r = radius from the earth’s center of mass; length unit r = distance from the earth’s center of mass; length unit S = scalar part of the impulse in the rotating frame; length unit/time unit t = time V = gravity potential; (length unit)2=(time unit)2 x; y; z = Cartesian coordinates of r; length unit α = earth’s equatorial radius; length unit ∆v = delta-v (impulse per unit of mass); length unit/time unit = auxiliary symbol of the order of the perturbation; dimensionless p η = eccentricity function η = 1 − e2; dimensionless θ = argument of latitude θ = f + !; rad µ = earth’s gravitational parameter; (length unit)3=(time unit)2 ρ, σ = scalar indices representing the perturbed dynamics; dimensionless τ = time of perigee passage (orbital element); time unit Ω= right ascension of the ascending node (orbital element); rad ! = argument of the perigee (orbital element); rad ∗ = superindex. The indexed magnitude is valid under the averaging assumption 1 Introduction It is demonstrated that the impulse in a single orbit of an earth’s satellite can be computed analytically within a reasonable accuracy, in this way providing a vec- torial expression that is used to reveal the more important features of the earth’s satellite dynamics without need of resorting to Fourier series expansions or pertur- bation theory. The impulse is referred to a rotating frame, and is computed in the “averaging assumption”, that is, taking the osculating semi-major axis, eccentric- ity, and inclination constant in the short time interval in which the argument of the latitude of the satellite advances by 2π. In particular, by analyzing the impulse due to the perturbation that arises from the J2 term of the earth’s gravitational potential it is shown the existence of planar orbits, which may exist in the equatorial plane and also in the meridian planes (polar orbits), as well as orbits with fixed perigee at the critical inclination —all of them in agreement with well-known results. In addition, it is shown that the information provided by this impulse can be encapsulated in simple scalar indices. The evaluation of these kinds of indices can be used for the creation of argument of the perigee vs. inclination maps, which 2 provide valuable information on the perturbed dynamics. Furthermore, the proce- dure is not restricted either to the J2 perturbation or to conservative perturbations, and can be generally applied, although in most cases the quadratures defining the impulse should be evaluated using numerical integration techniques. The impulse is computed per unit of mass, which therefore will produce a cor- responding variation of the velocity vector instead of the linear momentum. Hence, it provides the equivalent delta-v per orbit of gravitational perturbations, which can be useful in practical engineering problems. For instance, some maneuvers are de- signed using just the Keplerian approximation [1]. Also, the initial design of con- stellations is commonly made under the linear effect of the J2 perturbation, which does not include the effects of the perigee dynamics [2]. In both cases a better es- timation of the delta-v budget necessary to compensate the neglected effects may be obtained from the theory provided here. 2 ∆v on a Keplerian orbit The net delta-v from Keplerian motion between two times is Z t2 µ ∆v = − 3 r dt; (1) t1 r where the time dependence of r must be made explicit to perform the quadrature. This can be done using the orbital elements representation of r. x = cos Ω cos θ − sin Ω sin θ cos I; (2) r y = sin Ω cos θ + cos Ω sin θ cos I; (3) r z = sin θ sin I; (4) r where p r = ; (5) 1 + e cos f Equations (2)–(4) and (5) reveal the time dependence of the Cartesian coordi- nates of the Kepler problem, but this is done in an implicit way through the true anomaly. Although the true anomaly is an implicit function of time, the quadrature in Eq. (1) can be solved directly either in f or θ making use of the preservation of the total angular momentum of the Keplerian motion, which can be stated in the usual form dθ p r2 = kGk = µ p = na2η: (6) dt 3 where dt = (1=n) dM, and hence r2 n dt = dM = dθ; (7) a2η making in this way available an explicit relation between the differentials of the mean anomaly and the argument of the latitude. An analytical expression of the ∆v of Keplerian motion between two arbitrary times is obtained substituting Eq. (7) into Eq. (1) to give na Z θ(t2) r ∆v = − dθ; (8) η θ(t1) r which is integrated replacing r=r by its components in Eqs. (2)–(4), namely 0 1 θ(t2) − sin θ cos Ω − cos I cos θ sin Ω na ∆v = B − sin θ sin Ω + cos I cos θ cos Ω C : (9) η @ A sin I cos θ θ(t1) As expected, Eq. (9) vanishes when it is evaluated along a full orbit θ(t2)−θ(t1) = 2π. 3 Effects of J2 −3 For the earth, it happens that J2 = O(10 ) whereas all other coefficients are of order 10−6. Then, the non-central potential ! µ µ α2 1 z2 V = − + J −1 + 3 ; (10) r r r2 2 2 r2 is quite representative of a wide class of non-resonant low earth orbits. The Newtonian equations of motion derived from Eq. (10) are written in the vectorial form " ! # d2r µ µ α2 5 z2 1 = − r + 3 J − r − zk ; (11) dt2 r3 r3 r2 2 2 r2 2 whose solutions are no longer ellipses. However, as far as J2 is small, solutions to Eq. (11) may be viewed as slightly distorted ellipses whose deformation evolves with time, a case in which the concept of osculating elements can be used in the description of the non-Keplerian motion. Thus, the position and velocity of the body at a given moment in time can be used to define an instantaneous ellipse that 4 is tangent (osculating) to the actual trajectory. Since each instantaneous ellipse is defined by corresponding Keplerian elements, the non-Keplerian trajectory can be described by the time evolution of these osculating elements, to whom all the relations of the elliptic motion apply. The osculating elements representation is quite useful to manifest the time de- pendency of Eq. (11), in this way enabling the computation of the corresponding ∆v under the averaging assumption. That is, the osculating elements a(t), e(t), I(t), Ω(t), !(t), and τ(t) are assumed to evolve slowly when compared to the fast evolution of either the mean anomaly, which is driven by the osculating mean mo- tion n(t). Since the rate of variation of the argument of the latitude will only differ from n in effects of the order of the perturbation (J2 in the present case), which are related to the non-vanishing motion of argument of the perigee, these slow varying elements are taken as constants in the computation of the quadratures defining the ∆v in the interval in which θ increases by 2π, that is, a full orbit. 3.1 The rotating frame Note, however, that while taking a, e, and I constant along a full orbit may be a reasonable approximation, the assumption that ! and Ω remain constant in the same short time interval must be further qualified. In particular, as far as the right ascension of the ascending node may have a slightly different value in the initial and final times, independently of the smallness of this a priori unknown variation, taking Ω constant is equivalent to say that the quadratures are solved in a rotating frame which is moving with the rotation rate of Ω, a case in which corresponding inertia terms should be added to obtain the ∆v. The acceleration in the rotating frame is " ! # µ µ α2 5 z2 1 r¨ = − r + 3 J − r − zk − 2N × r_ − N × (N × r); (12) r3 r3 r2 2 2 r2 2 where N = Nk is the rotation rate of the node, which is further assumed to remain constant in the averaging assumption, and overdots are used to emphasize that corresponding derivatives are measured in the rotating frame. Then, the ∆v in the rotating frame is written ∆v = J K + J D + J C + J c; (13) where the Keplerian part J K is formally the same as in Eq.
Recommended publications
  • Appendix a Orbits
    Appendix A Orbits As discussed in the Introduction, a good ¯rst approximation for satellite motion is obtained by assuming the spacecraft is a point mass or spherical body moving in the gravitational ¯eld of a spherical planet. This leads to the classical two-body problem. Since we use the term body to refer to a spacecraft of ¯nite size (as in rigid body), it may be more appropriate to call this the two-particle problem, but I will use the term two-body problem in its classical sense. The basic elements of orbital dynamics are captured in Kepler's three laws which he published in the 17th century. His laws were for the orbital motion of the planets about the Sun, but are also applicable to the motion of satellites about planets. The three laws are: 1. The orbit of each planet is an ellipse with the Sun at one focus. 2. The line joining the planet to the Sun sweeps out equal areas in equal times. 3. The square of the period of a planet is proportional to the cube of its mean distance to the sun. The ¯rst law applies to most spacecraft, but it is also possible for spacecraft to travel in parabolic and hyperbolic orbits, in which case the period is in¯nite and the 3rd law does not apply. However, the 2nd law applies to all two-body motion. Newton's 2nd law and his law of universal gravitation provide the tools for generalizing Kepler's laws to non-elliptical orbits, as well as for proving Kepler's laws.
    [Show full text]
  • Astrodynamics
    Politecnico di Torino SEEDS SpacE Exploration and Development Systems Astrodynamics II Edition 2006 - 07 - Ver. 2.0.1 Author: Guido Colasurdo Dipartimento di Energetica Teacher: Giulio Avanzini Dipartimento di Ingegneria Aeronautica e Spaziale e-mail: [email protected] Contents 1 Two–Body Orbital Mechanics 1 1.1 BirthofAstrodynamics: Kepler’sLaws. ......... 1 1.2 Newton’sLawsofMotion ............................ ... 2 1.3 Newton’s Law of Universal Gravitation . ......... 3 1.4 The n–BodyProblem ................................. 4 1.5 Equation of Motion in the Two-Body Problem . ....... 5 1.6 PotentialEnergy ................................. ... 6 1.7 ConstantsoftheMotion . .. .. .. .. .. .. .. .. .... 7 1.8 TrajectoryEquation .............................. .... 8 1.9 ConicSections ................................... 8 1.10 Relating Energy and Semi-major Axis . ........ 9 2 Two-Dimensional Analysis of Motion 11 2.1 ReferenceFrames................................. 11 2.2 Velocity and acceleration components . ......... 12 2.3 First-Order Scalar Equations of Motion . ......... 12 2.4 PerifocalReferenceFrame . ...... 13 2.5 FlightPathAngle ................................. 14 2.6 EllipticalOrbits................................ ..... 15 2.6.1 Geometry of an Elliptical Orbit . ..... 15 2.6.2 Period of an Elliptical Orbit . ..... 16 2.7 Time–of–Flight on the Elliptical Orbit . .......... 16 2.8 Extensiontohyperbolaandparabola. ........ 18 2.9 Circular and Escape Velocity, Hyperbolic Excess Speed . .............. 18 2.10 CosmicVelocities
    [Show full text]
  • Flight and Orbital Mechanics
    Flight and Orbital Mechanics Lecture slides Challenge the future 1 Flight and Orbital Mechanics AE2-104, lecture hours 21-24: Interplanetary flight Ron Noomen October 25, 2012 AE2104 Flight and Orbital Mechanics 1 | Example: Galileo VEEGA trajectory Questions: • what is the purpose of this mission? • what propulsion technique(s) are used? • why this Venus- Earth-Earth sequence? • …. [NASA, 2010] AE2104 Flight and Orbital Mechanics 2 | Overview • Solar System • Hohmann transfer orbits • Synodic period • Launch, arrival dates • Fast transfer orbits • Round trip travel times • Gravity Assists AE2104 Flight and Orbital Mechanics 3 | Learning goals The student should be able to: • describe and explain the concept of an interplanetary transfer, including that of patched conics; • compute the main parameters of a Hohmann transfer between arbitrary planets (including the required ΔV); • compute the main parameters of a fast transfer between arbitrary planets (including the required ΔV); • derive the equation for the synodic period of an arbitrary pair of planets, and compute its numerical value; • derive the equations for launch and arrival epochs, for a Hohmann transfer between arbitrary planets; • derive the equations for the length of the main mission phases of a round trip mission, using Hohmann transfers; and • describe the mechanics of a Gravity Assist, and compute the changes in velocity and energy. Lecture material: • these slides (incl. footnotes) AE2104 Flight and Orbital Mechanics 4 | Introduction The Solar System (not to scale): [Aerospace
    [Show full text]
  • Elliptical Orbits
    1 Ellipse-geometry 1.1 Parameterization • Functional characterization:(a: semi major axis, b ≤ a: semi minor axis) x2 y 2 b p + = 1 ⇐⇒ y(x) = · ± a2 − x2 (1) a b a • Parameterization in cartesian coordinates, which follows directly from Eq. (1): x a · cos t = with 0 ≤ t < 2π (2) y b · sin t – The origin (0, 0) is the center of the ellipse and the auxilliary circle with radius a. √ – The focal points are located at (±a · e, 0) with the eccentricity e = a2 − b2/a. • Parameterization in polar coordinates:(p: parameter, 0 ≤ < 1: eccentricity) p r(ϕ) = (3) 1 + e cos ϕ – The origin (0, 0) is the right focal point of the ellipse. – The major axis is given by 2a = r(0) − r(π), thus a = p/(1 − e2), the center is therefore at − pe/(1 − e2), 0. – ϕ = 0 corresponds to the periapsis (the point closest to the focal point; which is also called perigee/perihelion/periastron in case of an orbit around the Earth/sun/star). The relation between t and ϕ of the parameterizations in Eqs. (2) and (3) is the following: t r1 − e ϕ tan = · tan (4) 2 1 + e 2 1.2 Area of an elliptic sector As an ellipse is a circle with radius a scaled by a factor b/a in y-direction (Eq. 1), the area of an elliptic sector PFS (Fig. ??) is just this fraction of the area PFQ in the auxiliary circle. b t 2 1 APFS = · · πa − · ae · a sin t a 2π 2 (5) 1 = (t − e sin t) · a b 2 The area of the full ellipse (t = 2π) is then, of course, Aellipse = π a b (6) Figure 1: Ellipse and auxilliary circle.
    [Show full text]
  • Perturbation Theory in Celestial Mechanics
    Perturbation Theory in Celestial Mechanics Alessandra Celletti Dipartimento di Matematica Universit`adi Roma Tor Vergata Via della Ricerca Scientifica 1, I-00133 Roma (Italy) ([email protected]) December 8, 2007 Contents 1 Glossary 2 2 Definition 2 3 Introduction 2 4 Classical perturbation theory 4 4.1 The classical theory . 4 4.2 The precession of the perihelion of Mercury . 6 4.2.1 Delaunay action–angle variables . 6 4.2.2 The restricted, planar, circular, three–body problem . 7 4.2.3 Expansion of the perturbing function . 7 4.2.4 Computation of the precession of the perihelion . 8 5 Resonant perturbation theory 9 5.1 The resonant theory . 9 5.2 Three–body resonance . 10 5.3 Degenerate perturbation theory . 11 5.4 The precession of the equinoxes . 12 6 Invariant tori 14 6.1 Invariant KAM surfaces . 14 6.2 Rotational tori for the spin–orbit problem . 15 6.3 Librational tori for the spin–orbit problem . 16 6.4 Rotational tori for the restricted three–body problem . 17 6.5 Planetary problem . 18 7 Periodic orbits 18 7.1 Construction of periodic orbits . 18 7.2 The libration in longitude of the Moon . 20 1 8 Future directions 20 9 Bibliography 21 9.1 Books and Reviews . 21 9.2 Primary Literature . 22 1 Glossary KAM theory: it provides the persistence of quasi–periodic motions under a small perturbation of an integrable system. KAM theory can be applied under quite general assumptions, i.e. a non– degeneracy of the integrable system and a diophantine condition of the frequency of motion.
    [Show full text]
  • New Closed-Form Solutions for Optimal Impulsive Control of Spacecraft Relative Motion
    New Closed-Form Solutions for Optimal Impulsive Control of Spacecraft Relative Motion Michelle Chernick∗ and Simone D'Amicoy Aeronautics and Astronautics, Stanford University, Stanford, California, 94305, USA This paper addresses the fuel-optimal guidance and control of the relative motion for formation-flying and rendezvous using impulsive maneuvers. To meet the requirements of future multi-satellite missions, closed-form solutions of the inverse relative dynamics are sought in arbitrary orbits. Time constraints dictated by mission operations and relevant perturbations acting on the formation are taken into account by splitting the optimal recon- figuration in a guidance (long-term) and control (short-term) layer. Both problems are cast in relative orbit element space which allows the simple inclusion of secular and long-periodic perturbations through a state transition matrix and the translation of the fuel-optimal optimization into a minimum-length path-planning problem. Due to the proper choice of state variables, both guidance and control problems can be solved (semi-)analytically leading to optimal, predictable maneuvering schemes for simple on-board implementation. Besides generalizing previous work, this paper finds four new in-plane and out-of-plane (semi-)analytical solutions to the optimal control problem in the cases of unperturbed ec- centric and perturbed near-circular orbits. A general delta-v lower bound is formulated which provides insight into the optimality of the control solutions, and a strong analogy between elliptic Hohmann transfers and formation-flying control is established. Finally, the functionality, performance, and benefits of the new impulsive maneuvering schemes are rigorously assessed through numerical integration of the equations of motion and a systematic comparison with primer vector optimal control.
    [Show full text]
  • Coordinate Systems in Geodesy
    COORDINATE SYSTEMS IN GEODESY E. J. KRAKIWSKY D. E. WELLS May 1971 TECHNICALLECTURE NOTES REPORT NO.NO. 21716 COORDINATE SYSTElVIS IN GEODESY E.J. Krakiwsky D.E. \Vells Department of Geodesy and Geomatics Engineering University of New Brunswick P.O. Box 4400 Fredericton, N .B. Canada E3B 5A3 May 1971 Latest Reprinting January 1998 PREFACE In order to make our extensive series of lecture notes more readily available, we have scanned the old master copies and produced electronic versions in Portable Document Format. The quality of the images varies depending on the quality of the originals. The images have not been converted to searchable text. TABLE OF CONTENTS page LIST OF ILLUSTRATIONS iv LIST OF TABLES . vi l. INTRODUCTION l 1.1 Poles~ Planes and -~es 4 1.2 Universal and Sidereal Time 6 1.3 Coordinate Systems in Geodesy . 7 2. TERRESTRIAL COORDINATE SYSTEMS 9 2.1 Terrestrial Geocentric Systems • . 9 2.1.1 Polar Motion and Irregular Rotation of the Earth • . • • . • • • • . 10 2.1.2 Average and Instantaneous Terrestrial Systems • 12 2.1. 3 Geodetic Systems • • • • • • • • • • . 1 17 2.2 Relationship between Cartesian and Curvilinear Coordinates • • • • • • • . • • 19 2.2.1 Cartesian and Curvilinear Coordinates of a Point on the Reference Ellipsoid • • • • • 19 2.2.2 The Position Vector in Terms of the Geodetic Latitude • • • • • • • • • • • • • • • • • • • 22 2.2.3 Th~ Position Vector in Terms of the Geocentric and Reduced Latitudes . • • • • • • • • • • • 27 2.2.4 Relationships between Geodetic, Geocentric and Reduced Latitudes • . • • • • • • • • • • 28 2.2.5 The Position Vector of a Point Above the Reference Ellipsoid . • • . • • • • • • . .• 28 2.2.6 Transformation from Average Terrestrial Cartesian to Geodetic Coordinates • 31 2.3 Geodetic Datums 33 2.3.1 Datum Position Parameters .
    [Show full text]
  • Mercury's Resonant Rotation from Secular Orbital Elements
    View metadata, citation and similar papers at core.ac.uk brought to you by CORE provided by Institute of Transport Research:Publications Mercury’s resonant rotation from secular orbital elements Alexander Stark a, Jürgen Oberst a,b, Hauke Hussmann a a German Aerospace Center, Institute of Planetary Research, D-12489 Berlin, Germany b Moscow State University for Geodesy and Cartography, RU-105064 Moscow, Russia The final publication is available at Springer via http://dx.doi.org/10.1007/s10569-015-9633-4. Abstract We used recently produced Solar System ephemerides, which incorpo- rate two years of ranging observations to the MESSENGER spacecraft, to extract the secular orbital elements for Mercury and associated uncer- tainties. As Mercury is in a stable 3:2 spin-orbit resonance these values constitute an important reference for the planet’s measured rotational pa- rameters, which in turn strongly bear on physical interpretation of Mer- cury’s interior structure. In particular, we derive a mean orbital period of (87.96934962 ± 0.00000037) days and (assuming a perfect resonance) a spin rate of (6.138506839 ± 0.000000028) ◦/day. The difference between this ro- tation rate and the currently adopted rotation rate (Archinal et al., 2011) corresponds to a longitudinal displacement of approx. 67 m per year at the equator. Moreover, we present a basic approach for the calculation of the orientation of the instantaneous Laplace and Cassini planes of Mercury. The analysis allows us to assess the uncertainties in physical parameters of the planet, when derived from observations of Mercury’s rotation. 1 1 Introduction Mercury’s orbit is not inertially stable but exposed to various perturbations which over long time scales lead to a chaotic motion (Laskar, 1989).
    [Show full text]
  • SATELLITES ORBIT ELEMENTS : EPHEMERIS, Keplerian ELEMENTS, STATE VECTORS
    www.myreaders.info www.myreaders.info Return to Website SATELLITES ORBIT ELEMENTS : EPHEMERIS, Keplerian ELEMENTS, STATE VECTORS RC Chakraborty (Retd), Former Director, DRDO, Delhi & Visiting Professor, JUET, Guna, www.myreaders.info, [email protected], www.myreaders.info/html/orbital_mechanics.html, Revised Dec. 16, 2015 (This is Sec. 5, pp 164 - 192, of Orbital Mechanics - Model & Simulation Software (OM-MSS), Sec 1 to 10, pp 1 - 402.) OM-MSS Page 164 OM-MSS Section - 5 -------------------------------------------------------------------------------------------------------43 www.myreaders.info SATELLITES ORBIT ELEMENTS : EPHEMERIS, Keplerian ELEMENTS, STATE VECTORS Satellite Ephemeris is Expressed either by 'Keplerian elements' or by 'State Vectors', that uniquely identify a specific orbit. A satellite is an object that moves around a larger object. Thousands of Satellites launched into orbit around Earth. First, look into the Preliminaries about 'Satellite Orbit', before moving to Satellite Ephemeris data and conversion utilities of the OM-MSS software. (a) Satellite : An artificial object, intentionally placed into orbit. Thousands of Satellites have been launched into orbit around Earth. A few Satellites called Space Probes have been placed into orbit around Moon, Mercury, Venus, Mars, Jupiter, Saturn, etc. The Motion of a Satellite is a direct consequence of the Gravity of a body (earth), around which the satellite travels without any propulsion. The Moon is the Earth's only natural Satellite, moves around Earth in the same kind of orbit. (b) Earth Gravity and Satellite Motion : As satellite move around Earth, it is pulled in by the gravitational force (centripetal) of the Earth. Contrary to this pull, the rotating motion of satellite around Earth has an associated force (centrifugal) which pushes it away from the Earth.
    [Show full text]
  • Impulsive Maneuvers for Formation Reconfiguration Using Relative Orbital Elements
    JOURNAL OF GUIDANCE,CONTROL, AND DYNAMICS Vol. 38, No. 6, June 2015 Impulsive Maneuvers for Formation Reconfiguration Using Relative Orbital Elements G. Gaias∗ and S. D’Amico† DLR, German Aerospace Center, 82234 Wessling, Germany DOI: 10.2514/1.G000189 Advanced multisatellite missions based on formation-flying and on-orbit servicing concepts require the capability to arbitrarily reconfigure the relative motion in an autonomous, fuel efficient, and flexible manner. Realistic flight scenarios impose maneuvering time constraints driven by the satellite bus, by the payload, or by collision avoidance needs. In addition, mission control center planning and operations tasks demand determinism and predictability of the propulsion system activities. Based on these considerations and on the experience gained from the most recent autonomous formation-flying demonstrations in near-circular orbit, this paper addresses and reviews multi- impulsive solution schemes for formation reconfiguration in the relative orbit elements space. In contrast to the available literature, which focuses on case-by-case or problem-specific solutions, this work seeks the systematic search and characterization of impulsive maneuvers of operational relevance. The inversion of the equations of relative motion parameterized using relative orbital elements enables the straightforward computation of analytical or numerical solutions and provides direct insight into the delta-v cost and the most convenient maneuver locations. The resulting general methodology is not only able to refind and requalify all particular solutions known in literature or flown in space, but enables the identification of novel fuel-efficient maneuvering schemes for future onboard implementation. Nomenclature missions. Realistic operational scenarios ask for accomplishing such a = semimajor axis actions in a safe way, within certain levels of accuracy, and in a fuel- B = control input matrix of the relative dynamics efficient manner.
    [Show full text]
  • 2. Orbital Mechanics MAE 342 2016
    2/12/20 Orbital Mechanics Space System Design, MAE 342, Princeton University Robert Stengel Conic section orbits Equations of motion Momentum and energy Kepler’s Equation Position and velocity in orbit Copyright 2016 by Robert Stengel. All rights reserved. For educational use only. http://www.princeton.edu/~stengel/MAE342.html 1 1 Orbits 101 Satellites Escape and Capture (Comets, Meteorites) 2 2 1 2/12/20 Two-Body Orbits are Conic Sections 3 3 Classical Orbital Elements Dimension and Time a : Semi-major axis e : Eccentricity t p : Time of perigee passage Orientation Ω :Longitude of the Ascending/Descending Node i : Inclination of the Orbital Plane ω: Argument of Perigee 4 4 2 2/12/20 Orientation of an Elliptical Orbit First Point of Aries 5 5 Orbits 102 (2-Body Problem) • e.g., – Sun and Earth or – Earth and Moon or – Earth and Satellite • Circular orbit: radius and velocity are constant • Low Earth orbit: 17,000 mph = 24,000 ft/s = 7.3 km/s • Super-circular velocities – Earth to Moon: 24,550 mph = 36,000 ft/s = 11.1 km/s – Escape: 25,000 mph = 36,600 ft/s = 11.3 km/s • Near escape velocity, small changes have huge influence on apogee 6 6 3 2/12/20 Newton’s 2nd Law § Particle of fixed mass (also called a point mass) acted upon by a force changes velocity with § acceleration proportional to and in direction of force § Inertial reference frame § Ratio of force to acceleration is the mass of the particle: F = m a d dv(t) ⎣⎡mv(t)⎦⎤ = m = ma(t) = F ⎡ ⎤ dt dt vx (t) ⎡ f ⎤ ⎢ ⎥ x ⎡ ⎤ d ⎢ ⎥ fx f ⎢ ⎥ m ⎢ vy (t) ⎥ = ⎢ y ⎥ F = fy = force vector dt
    [Show full text]
  • Orbital Mechanics Course Notes
    Orbital Mechanics Course Notes David J. Westpfahl Professor of Astrophysics, New Mexico Institute of Mining and Technology March 31, 2011 2 These are notes for a course in orbital mechanics catalogued as Aerospace Engineering 313 at New Mexico Tech and Aerospace Engineering 362 at New Mexico State University. This course uses the text “Fundamentals of Astrodynamics” by R.R. Bate, D. D. Muller, and J. E. White, published by Dover Publications, New York, copyright 1971. The notes do not follow the book exclusively. Additional material is included when I believe that it is needed for clarity, understanding, historical perspective, or personal whim. We will cover the material recommended by the authors for a one-semester course: all of Chapter 1, sections 2.1 to 2.7 and 2.13 to 2.15 of Chapter 2, all of Chapter 3, sections 4.1 to 4.5 of Chapter 4, and as much of Chapters 6, 7, and 8 as time allows. Purpose The purpose of this course is to provide an introduction to orbital me- chanics. Students who complete the course successfully will be prepared to participate in basic space mission planning. By basic mission planning I mean the planning done with closed-form calculations and a calculator. Stu- dents will have to master additional material on numerical orbit calculation before they will be able to participate in detailed mission planning. There is a lot of unfamiliar material to be mastered in this course. This is one field of human endeavor where engineering meets astronomy and ce- lestial mechanics, two fields not usually included in an engineering curricu- lum.
    [Show full text]